You are on page 1of 19

NIH Public Access

Author Manuscript
Life Sci. Author manuscript; available in PMC 2015 February 27.
Published in final edited form as:
NIH-PA Author Manuscript

Life Sci. 2014 February 27; 97(1): 55–63. doi:10.1016/j.lfs.2013.09.011.

Moving around the molecule: Relationship between chemical


structure and in vivo activity of synthetic cannabinoids
Jenny L. Wileya, Julie A. Marusicha, and John W. Huffmanb
aRTI International, 3040 Cornwallis Road, Research Triangle Park, NC

bDepartment of Chemistry, Clemson University, Clemson, SC

Abstract
Originally synthesized for research purposes, indole- and pyrrole-derived synthetic cannabinoids
are the most common psychoactive compounds contained in abused products marketed as “spice”
or “herbal incense.” While CB1 and CB2 receptor affinities are available for most of these research
chemicals, in vivo pharmacological data are sparse. In mice, cannabinoids produce a characteristic
NIH-PA Author Manuscript

profile of dose-dependent effects: antinociception, hypothermia, catalepsy and suppression of


locomotion. In combination with receptor binding data, this tetrad battery has been useful in
evaluation of the relationship between the structural features of synthetic cannabinoids and their in
vivo cannabimimetic activity. Here, published tetrad studies are reviewed and additional in vivo
data on synthetic cannabinoids are presented. Overall, the best predictor of likely cannabimimetic
effects in the tetrad tests was good CB1 receptor affinity. Further, retention of good CB1 affinity
and in vivo activity was observed across a wide array of structural manipulations of substituents of
the prototypic aminoalkylindole molecule WIN55,212-2, including substitution of an alkyl for the
morpholino group, replacement of an indole core with a pyrrole or phenylpyrrole, substitution of a
phenylacetyl or tetramethylcyclopropyl group for JWH-018’s naphthoyl, and halogenation of the
naphthoyl group. This flexibility of cannabinoid ligand-receptor interactions has been a particular
challenge for forensic scientists who have struggled to identify and regulate each new compound
as it has appeared on the drug market. One of the most pressing future research needs is
determination of the extent to which the pharmacology of these synthetic cannabinoids may differ
from those of classical cannabinoids.

Keywords
NIH-PA Author Manuscript

alkylindoles; aromatic stacking; cannabinoids; herbal marijuana; indoles; JWH-018; pyrroles;


review; spice; synthetic cannabinoids

© 2013 Elsevier Inc. All rights reserved.


Corresponding Author: Jenny Wiley, Ph.D., RTI International, 3040 Cornwallis Road, Research Triangle Park, North Carolina
27709-2194, Phone: (919) 541-7276, FAX: (919) 541-6499, jwiley@rti.org.
Conflict of interest statement
The authors do not have any real or perceived conflicts of interest with regard to the data presented in this manuscript.
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our
customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of
the resulting proof before it is published in its final citable form. Please note that during the production process errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Wiley et al. Page 2

Introduction
The use of marijuana and other constituents of the cannabis plant (e.g., hashish) for
NIH-PA Author Manuscript

medicinal and religious purposes can be traced back to ancient China. Yet, intensive
scientific interest in marijuana awaited development of appropriate tools, the first of which
arrived in 1964 when Dr. Raphael Mechoulam and colleagues isolated and identified Δ9-
tetrahydrocannabinol (Δ9-THC) as the primary psychoactive substituent of the marijuana
plant (Gaoni and Mechoulam, 1964). Continuing his work in the emerging field of
cannabinoid science, Professor Mechoulam founded a lab that served as an incubator for
several foundational discoveries that occurred later in the 1980’s and 1990’s. During these
formative years, major research findings included discovery of the endocannabinoid system
(Devane, et al., 1992; Hanus, et al., 2001; Mechoulam, et al., 1995), identification and
cloning of the CB1 and CB2 subtypes of cannabinoid receptor (Devane, et al., 1988;
Matsuda, et al., 1990; Munro, et al., 1993), and synthesis of selective antagonists for each of
these receptors (Rinaldi-Carmona, et al., 1994; Rinaldi-Carmona, et al., 1998). Although
cannabinoid agonists used in initial in vivo research studies were phytocannabinoids such as
Δ9-THC and cannabidiol (Compton, et al., 1990; Martin, et al., 1981), these plant-derived
cannabinoids soon shared the research arena with synthetic cannabinoids, including the
novel classes of bicyclic cannabinoids (e.g., CP55,940; Little, et al., 1988) and the
aminoalkylindoles (e.g., WIN55,212-2; Compton, et al., 1992), developed by scientists at
Pfizer and Sterling-Winthrop, respectively, as well as the endocannabinoids (e.g.,
NIH-PA Author Manuscript

anandamide and 2-arachidonoylglycerol; Devane, et al., 1992; Mechoulam, et al., 1995).

The synthetic cannabinoids, which were later hijacked for use in designer drugs labeled as
“herbal incense,” were originally synthesized in this milieu (Huffman, et al., 1994).
JWH-018 [1-pentyl-3-(1-naphthoyl)indole], one of the first such compound identified in a
forensic sample (Lindigkeit, et al., 2009), and other structurally related synthetic
cannabinoids were created as part of a research program directed by Dr. John Huffman at
Clemson University, Clemson, South Carolina. In effort to determine how such structurally
diverse molecules as Δ9-THC and WIN55,212-2 could fit into the same receptor, Dr.
Huffman considered the then popular 3-point attachment model of cannabinoid ligand-
receptor interaction (Thomas, et al., 1991) and hypothesized that the attachment location for
the morpholino group of WIN would correspond with the pentyl side chain of Δ9-THC
(Figure 1). To test this hypothesis, he synthesized a series of indole- and pyrrole-derived
cannabinoids and elicited the assistance of Dr. Billy Martin and colleagues at Virginia
Commonwealth University, Richmond, Virginia to evaluate them. Dr. Martin’s structure-
activity relationship (SAR) analysis of synthetic cannabinoids used a two-step approach to
measure binding affinity at CB1 and CB2 receptors and to assess the cannabinoids in a tetrad
of in vivo tests in which cannabinoid agonists produce a characteristic profile of effects in
NIH-PA Author Manuscript

mice, including suppression of motor activity, antinociception, hypothermia and catalepsy


(Martin, et al., 1991). This mini-review focuses on summarizing the findings of published
SAR studies which examined the in vivo activity of indole- and pyrrole-derived
cannabinoids. In addition, new in vivo data are presented for structurally related indoles and
pyrroles. These unpublished data were collected in the same lab and under the same
experimental parameters used for the previously published data (Wiley, et al., 2012a; Wiley,
et al., 2012b; Wiley, et al., 1998). Animals used in these experiments were cared for in
accordance with the guidelines of the Institutional Animal Care and Use Committee of the
Virginia Commonwealth University (Richmond, Virginia) and the ‘Guide For The Care And
Use Of Animals’ (National Research Council, 1996). Not covered in this review are
manuscripts that present only binding and/or functional in vitro data. The interested reader is
referred to several excellent reviews that have been written on this topic (Huffman, 1999;
Huffman, 2009; Huffman and Padgett, 2005; Manera, et al., 2008). In addition, other
manuscripts in this volume review separate aspects of the in vivo and behavioral

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 3

pharmacology of this class of cannabinoids, including their discriminative stimulus effects


(Järbe and Gifford) and their liability for producing tolerance and dependence (Lichtman et
al..).
NIH-PA Author Manuscript

WIN55,212-2
WIN55,212-2 represents the prototypic aminoalkylindole cannabinoid which inspired the
idea for synthetic indole cannabinoids (Huffman, et al., 1994). Coincidentally, the
enantiomers of WIN55,212 and Δ9-THC with cannabinoid activity are polar opposites: (+)-
WIN55,212 (designated WIN55,212-2) and (−)-Δ9-THC. (+)-Δ9-THC and the (−)-
enantiomer, (−)-WIN55,212 (designated as WIN55,212-3) are not active in the tetrad test
battery in mice (Compton, et al., 1992; Martin, et al., 1991). While Δ9-THC binds with
approximately equal affinity to both identified cannabinoid receptors (CB1 Ki = 41 nM, CB2
Ki = 36), WIN55,212-2 has better affinity for CB2 (Ki = 0.28 nM) vs CB1 (Ki = 1.89 nM)
receptors (Showalter, et al., 1996). Δ9-THC and WIN55,212-2 also differ in their in vitro
efficacy at the CB1 receptor, with Δ9-THC acting as a partial agonist in functional assays
such as [35S]GTPγS binding and WIN55,212-2 acting as a full agonist (Breivogel and
Childers, 2000); however, both compounds show approximately equal efficacy in the mouse
tetrad tests. In the present study, a full dose-effect curve determination was conducted with
WIN55,212-2 in the tetrad battery (Figure 2) and serves as a comparison for other novel
synthetic indole and pyrrole cannabinoids shown in Tables 1–4. Results are similar to those
NIH-PA Author Manuscript

that have been obtained previously with WIN55,212-2 (Compton, et al., 1992), Δ9-THC
(Martin, et al., 1991), and a variety of other psychoactive cannabinoid agonists (Compton, et
al., 1993). At lower doses, WIN55,212-2 (present study; Figure 2, panel A) may increase
locomotion, whereas at higher doses cannabinoid agonists, including WIN55,212-2 (present
study; Compton, et al., 1992), reliably and dose-dependently suppress locomotor activity
(Martin, et al., 1991). Maximum suppression typically approaches 100%. WIN55,212-2 also
produced robust antinociceptive effects (100% maximum possible effect at higher doses;
Figure 2, panel B) and substantial decreases in rectal temperature (−6 °C is typical at higher
doses; Figure 2, panel C). WIN55,212-2 increases catalepsy in the mouse ring test, with
efficacy ranging from 60 – 100% (Figure 2, panel D). Acute Δ9-THC produces similar
biphasic locomotor effects, decreases in rectal temperature, and increases in catalepsy
(Martin, et al., 1991; Sañudo-Peña, et al., 2000; Wiley, et al., 2008). ED50s for the effects of
WIN55,212-2 in these tests are provided in Table 1.

Indoles
Substitutions for morpholino group
The initial series of indole-derived cannabinoids that were used to evaluate the hypothesis
NIH-PA Author Manuscript

that the morpholino group of the aminoalkylindoles and the C3 side chain of Δ9-THC
occupied similar spatial domains within the CB1 receptor were a series of indole compounds
in which an alkyl group was substituted for the oxazine and morpholino substituents of
WIN55,212-2. One of the primary findings of this study was that a cyclic amino group at
this position was unnecessary for psychoactivity. Substitution of a n-pentyl group resulted in
a compound (JWH-018; Figure 1) with good CB1 receptor affinity (Ki = 9 nM) and
correspondingly good potency in the tetrad tests (Wiley, et al., 1998). Further, affinity and
potency varied systematically with the length of the carbon chain (position R1 of
naphthoylindole structure in Figure 3), with optimal activity from n-butyl to n-hexyl and
absence or reduction of receptor binding at shorter or longer carbon chains. This finding is
consistent with the results of a structure-activity analysis of classical and bicyclic
cannabinoids, in which length and branching of the C3 carbon chain of Δ8-THC and
CP55,940 analogs dramatically affected CB1 affinity and in vivo potency (Compton, et al.,
1993). Fluoropentyl substitution also resulted in an active compound (AM-2201) that has

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 4

been found in “herbal incense” products (Denooz, et al., 2013; Logan, et al., 2012). CB1
receptor recognition and in vivo activity were also retained with some cyclic substitutions
(cyclohexylethyl and cyclopropylmethyl), but not with others (2-phenylethyl), albeit none of
NIH-PA Author Manuscript

these compounds were as potent as WIN55,212-2 or JWH-018 (Wiley, et al., 1998). Of the
compounds tested in this early study, the most potent compounds (JWH-073, JWH-018,
JWH-019, and AM-2201: n-butyl, n-pentyl, n-hexyl, and n-fluorpentyl substitutions,
respectively) were detected in samples purchased in 2011 in the state of North Carolina in
the United States (Cox, et al., 2012), suggesting that manufacturers of these now illicit
products may have been aware of these initial findings.

Addition of substituents to indole group


The n-alkyl indoles described above were unmethylated at the 2-indole position (position R2
of naphthoylindole structure in Figure 3). A parallel set synthesized during the same time
period contained a 2-methylindole substituent. While 2-methylation did not affect the
overall pattern of results (i.e., optimal affinity and potency with n-butyl to n-hexyl chain
lengths), absolute affinities and potencies the methylated n-butyl to n-hexyl analogs were 2-
to 4-fold less than those of the unmethylated compounds (Wiley, et al., 1998). Further, 2-
methylation appeared to result in a shift in the SAR profile such that the methylated n-propyl
was more active and the n-heptyl compound was less active than the corresponding
unmethylated compounds. These results are consistent with those described for the
aminoalkylindole analogs of WIN55,212-2, in that the 2-methylated compound,
NIH-PA Author Manuscript

WIN53,365, showed 2-fold less affinity for CB1 receptors than did its unmethylated
counterpart, WIN55,225 (Dutta, et al., 1997).

The effects of a series of compounds with C-5 halogenation (Br and I) of the indole group
on binding and in vivo activity are presented in Table 1. Compared to compounds with n-
alkyl groups of comparable lengths, these halogenated indoles exhibited greatly decreased
affinity for the CB1 receptor. For example, JWH-455 (Ki = 686 nM) showed 77-fold less
CB1 receptor affinity than did JWH-073 (Ki = 8.9 nM), the non-halogenated n-butyl analog.
As with the non-halogenated series, optimal chain length was n-pentyl (Ki = 145 and 178 for
the Br analog JWH-444 and the I analog JWH-452, respectively). Unlike the original non-
halogenated compounds, however, none of the analogs in Table 1 showed < 100 nM CB1
receptor affinity. Consequently, a single 30 mg/kg intravenous dose of each compound was
assessed in vivo (where available supply allowed). Although one compound (JWH-453)
showed poor CB1 receptor affinity and was not active in any of the tests, all of the other
tested compounds exhibited activity or partial activity in one or more test. Given the poor
CB1 receptor binding affinity of some of the compounds (e.g., JWH-446 and JWH-455),
these results are somewhat surprising. One possibility is that these compounds may be
NIH-PA Author Manuscript

metabolized to active metabolites, as has been demonstrated for other indole-derived


cannabinoids (Brents et al., 2012). On the other hand, the tetrad tests are not entirely
selective for cannabinoids (Wiley and Martin, 2003), suggesting that non-cannabinoid
receptor actions may have played a role in the effects of these compounds at the 30 mg/kg
probe dose. Assessment of the ability of rimonabant to reverse the effects would be
necessary to determine whether the observed effects were CB1 receptor mediated.

Pyrroles
Naphthoylpyrroles
Conversion of the original series of naphthoylindoles to naphthoylpyrroles resulted in
compounds with greatly reduced CB1 receptor affinities and in vivo potencies (Wiley, et al.,
1998). Compounds with chain lengths of n-methyl to n-propyl (position R1 of
naphthoylpyrrole structure in Figure 3) did not bind to the CB1 receptor (Ki > 10,000 nM).

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 5

Whereas lengthening the chain to n-butyl or longer improved affinity, only the n-pentyl
pyrrole (JWH-030) had reasonable affinity (Ki = 87 nM), although its potency in at least two
of the in vivo tests (temperature and ring immobility) remained low (> 70 μM/kg). Perhaps
NIH-PA Author Manuscript

not surprisingly, these naphthoylpyrroles are not commonly detected in forensic samples,
unlike the comparable naphthoylindole analogs (e.g., JWH-018, JWH-073), albeit there are
exceptions: e.g., 3-naphthoyl-n-pentylpyrrole (JWH-030) has been detected in samples
obtained over the internet (Uchiyama, et al., 2013).

Phenylpyrroles
Although the 3-point attachment model formed the original basis of the creation of indole-
derived cannabinoids, exploration of their SAR led to revision of the model in favor of the
hypothesis that interaction of these new cannabinoids with the CB1 receptor might involve
stacking of their aromatic naphthoyl and indole substituents (Huffman, et al., 2003; Reggio,
et al., 1998). Later synthesis of a series of phenylpyrroles and evaluation of CB1 receptor
binding data was supportive of this aromatic stacking hypothesis (Huffman, et al., 2006).
Table 2 presents unpublished results of in vivo tetrad tests with some of these compounds.
Due to limited supplies, a full dose-effect curve could not be determined for all of the
compounds. Hence, potencies could not be estimated; however, several observations may be
made based upon these data. First, with the exception of JWH-309 (which could not be
tested at an adequate dose due to limited supply), all of the compounds were fully or
partially active in one or more of the tetrad tests. Second, these results are generally
NIH-PA Author Manuscript

consistent with the excellent CB1 receptor binding affinities of the majority of the
compounds. JWH-363 had the worst binding affinity (Ki=245 nM) and showed the least
activity in vivo (at 30 mg/kg i.v., inactive in two tests and minimal activity in the other two).
In contrast, JWH-370 had the best binding affinity (Ki=5.6 nM) and was fully active in all
four tests at a dose of 10 mg/kg, i.v. Third, both the nature and the position of the phenyl
substituent affected in vivo activity. For example, compounds with 2- or 3-phenyl
substituents (JWH-370, JWH-346, JWH-365, JWH-367, and JWH-307) were fully active in
the complete tetrad whereas compounds with 4-phenyl substituents (JWH-364 and
JWH-371) were inactive in at least one of the tests, despite reasonable CB1 binding
affinities. Trifluoromethylphenyl substitution (JWH-372 and JWH-363) resulted in
decreased activity in one or more of the tests, as compared to compounds with other 2- or 3-
phenyl substituents. Together, these results demonstrate overall concordance of in vivo
activity and good CB1 receptor binding affinities for the compounds and are supportive of
the aromatic stacking hypothesis.

Naphthoyl manipulations
Addition of substituents to naphthoyl group
NIH-PA Author Manuscript

The steric and electronic effects of two types of substituents on the naphthoyl group have
been examined in SAR with an in vivo component: electron withdrawing halogen
substituents and electron donating methoxy. The effects of moderately electron withdrawing
halogen substituents (Br, Cl, F, and I) at C-4 or C-8 (see chemical structure at the top of
Table 3) on CB1 and CB2 receptor affinities and in vivo pharmacology were examined in a
series of 1-alkyl-3-(1-naphthoyl)indoles (Wiley, et al., 2012b). In addition, two
electroneutral structural features (the length of the N-alkyl group and the presence of a
methyl at the 2-indole position) were manipulated in this series. Results observed with the
two electroneutral structural manipulations were consistent with a number of previous SAR
investigations of indole- and pyrrole-derived cannabinoids in which these structural
manipulations produce similar patterns of alterations of CB1 and CB2 affinities and in vivo
potencies, suggesting that the length of the alkyl substituent is as crucial for this cannabinoid
structural motif (Wiley, et al., 1998) as it is for cannabinoids based on the

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 6

tetrahydrocannabinol structural motif (Compton, et al., 1993). As found in these previous


studies, 1-pentyl substitution was optimal for enhancement of CB1 receptor affinity and in
vivo activity. Further, C-4 substituents showed accentuated CB1 receptor affinity and in vivo
NIH-PA Author Manuscript

activity compared to C-8 substituents, some of which did not bind to the receptor nor
produce activity in the tetrad tests. Less variability occurred across halogen groups. The
authors hypothesized that steric interference with aromatic stacking was a probable cause of
the decreased activity of compounds with C-8 substituents (Wiley, et al., 2012b).

The effects of methoxy substitution on the naphthoyl group appear to support this hypothesis
(Table 3; Aung, et al., 2000). Although halogen and methoxy substituents are electronic
opposites (electron withdrawing vs. electron donating, respectively), C-4 substitution of
either type of substituent resulted in compounds with the best CB1 receptor affinities and in
vivo activity, as compared to substitution at other positions. Unlike substituents at other
positions on the naphthoyl, the rotation of C-4 substituents is less hindered and thereby, less
likely to interfere with optimal aromatic stacking, which has been shown to be important for
cannabinoid receptor recognition (Huffman, et al., 2003; Reggio, et al., 1998). Consistent
with this idea, three of the four compounds (JWH-389 JWH-390, and JWH-391) with C-2
methoxy subsituents did not bind to the CB1 receptor. Further, none of the C-2 compounds
were active in any of the tetrad tests at doses up to 30 mg/kg i.v. (Table 3). While
compounds with C-6 methoxy substituents showed slightly increased activity across tests,
none of these compounds were fully active in all four tests. In addition, as in previous
NIH-PA Author Manuscript

studies, the length of the 1-alkyl and presence or absence of 2-methylation also contributed
to affinity for the cannabinoid receptors and activity in the tetrad tests (Table 3). Together,
these results suggest that steric effects play a stronger role in determining the nature of CB1
receptor affinity and in vivo activity than did electronic effects.

Substitution for naphthoyl group


Because the naphthoyl group itself is involved in aromatic stacking of indole- and pyrrole-
derived cannabinoids described thus far in this review, its manipulation would be expected
to produce profound changes in CB1 receptor recognition and in vivo activity and, indeed, it
does. Deletion of the naphthoyl (JWH-001) or substitution of a non-cyclic group (JWH-002)
eliminates both CB1 receptor binding affinity and in vivo activity (Table 4; Huffman, et al.,
1994). The number of aromatic constitutents also may play a role in interaction with the CB1
receptor and resultant pharmacological activity in the tetrad tests. For example, WIN56,098,
an aminoalkylindole analog with anthracene (3 aromatic ring) substitution for the
naphthalene, does not bind to the CB1 receptor nor is it active in vivo in the tetrad tests
(Compton, et al., 1992). In contrast, substitution of a two-ringed group (methyl-
benzodioxole) results in a 3,4-meylenedioxyphenylacetylindole (JWH-317) that binds to
CB1 receptors with moderate affinity (Ki=101 nM) and is active with near maximal efficacy
NIH-PA Author Manuscript

in suppressing locomotor activity and producing antinociception and hypothermia (Table 4).
Decreasing the number of aromatic rings on the non-indole side of the carbonyl to one, as in
a series of 1-pentyl-3-phenylacetylindoles (Figure 3), attenuated CB1 receptor affinities and
reduced in vivo potencies compared to 1-pentyl-3-naphthoylindole congeners (Wiley, et al.,
2012a). The major structural manipulations in this series included the type of substituent
(i.e., unsubstituted, methyl, methoxy, chloro, bromo, and fluoro) and the position of the
substituent on the phenyl ring (i.e., 2-, 3- or 4-position). Of these manipulations, the most
critical factor affecting in vivo potency was the position of the substituent. Whereas
compounds with 2- and 3-phenylacetyl substituents were efficacious with good potencies, 4-
substituents resulted in compounds that had poor potency or were inactive. Interestingly, this
pattern also occurred for indole-derived cannabinoids with a 1-morpholinomethtyl
substituent instead of a 1-pentyl (Compton, et al., 1992). For example, pravadoline
(WIN48,098), one of Sterling-Winthrop’s original lead compounds in the aminoalkylindole

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 7

series (Haubrich, et al., 1990), has a 4-methoxyphenyl substituent in place of the naphthoyl.
Pravadoline, and a similar analog without 2-methylation, do not bind to the CB1 receptor
and are not active in the tetrad tests (Compton, et al., 1992). This pattern of results suggests
NIH-PA Author Manuscript

that steric influences are as important for aminoalkylindoles and 1-pentyl-3-


phenylacetylindoles as they are for the halogenated 1-alkyl-3-(1-naphthoyl)indoles (Wiley,
et al., 2012b).

Recently, we completed assessment of two tetramethylcyclopropyl ketone indoles, UR-144


[(1-pentyl-1H-indol-3-yl)-(2,2,3,3-tetramethylcyclopropyl)methanone] and XLR-11 [(1-(5-
fluoropentyl)-1H-indol-3-yl)-(2,2,3,3-tetramethylcyclopropyl)methanone]. In these
compounds, a tetramethylcyclopropyl group is substituted for the naphthoyl of the parent 1-
pentyl-3-naphthoylindole (Figure 3). In addition, XLR-11 has a fluoro group at the terminal
end of the 1-pentyl (position R1 of tetramethylcyclopropyl ketone indole structure in Figure
3). These compounds are similar to those contained in a series of compounds synthesized by
Abbott Laboratories (Frost, et al., 2010; Frost, et al., 2008), suggesting that manufacturers of
“herbal incense” products have used multiple sources of information in their quest for novel
psychoactive cannabinoids. While the Abbott compounds were CB2 receptor selective,
many also possessed significant affinity for the CB1 receptor. As other compounds (e.g.,
JWH-018, AM-2201) have been banned, UR-144 and XLR-11 have started appearing in
samples of “herbal incense” sold over the internet (Kavanagh, et al., 2013; Uchiyama, et al.,
2013). Both compounds have good CB1 and CB2 receptor binding affinities and produced
NIH-PA Author Manuscript

the full complement of cannabimimetic effects in the tetrad tests in mice, with potencies
several-fold greater than Δ9-THC (Wiley et al., 2013). Given that attenuation of cannabinoid
activity was observed with the 1-pentyl-3-phenylacetylindoles (JWH-205 and JWH-167,
with and without 2-methylation, respectively), this result is rather surprising and suggests
that the nature of the cyclic substituent (e.g., phenyl vs. tetramethylcyclopropyl), as well as
their number, contribute to receptor recognition and in vivo potency. Both the
phenylacetylindoles and the Abbott compounds have shown up in “herbal incense” products.

Conclusions
Originally synthesized for research purposes, indole- and pyrrole-derived synthetic
cannabinoids are the most commonly identified psychoactive chemicals contained in
products marketed as “herbal incense.” Although hundreds of these cannabinoids have been
evaluated for their CB1 and CB2 receptor affinities (Aung, et al., 2000; Huffman, et al.,
2006; Huffman, et al., 2003; Huffman, et al., 2005a; Huffman, et al., 2005b; Lainton, et al.,
1995; Manera, et al., 2008), most have never been tested in animals before they were
discovered in products confiscated from human users. The purpose of this mini-review was
to summarize the findings of SAR studies in which the in vivo activity of these cannabinoids
NIH-PA Author Manuscript

was examined. Based upon the collective data, good CB1 affinity was the best predictor of
the propensity of a given compound to produce cannabimimetic effects in the tetrad tests in
mice, suggesting that these effects are mediated by CB1 receptor activation, as has also been
shown for traditional and bicyclic cannabinoids (Compton, et al., 1996; Compton, et al.,
1993). Consistent with this idea, reversal of psychoactive effects of indole-derived
cannabinoids has been demonstrated in nonhuman primates, rats and mice (Ginsburg, et al.,
2012; Järbe, et al., 2011; Wiebelhaus, et al., 2012). Another overall observation is that good
CB1 affinity and its associated cannabimimetic in vivo activity were retained across a
surprisingly wide array of structural manipulations of substituents of the prototypic
WIN55,212-2 molecule, including substitution of an alkyl for the WIN55,212-2 morpholino
group, replacement of an indole core with a pyrrole or phenylpyrrole, substitution of a
phenylacetyl or tetramethylcyclopropyl group for JWH-018’s naphthoyl, and halogenation
of the naphthoyl group. This flexibility of cannabinoid ligand-receptor interactions has
severely undermined the efforts of forensic scientists who have struggled to identify and

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 8

regulate each new compound as it has appeared in marketed products (Grabenauer, et al.,
2012; Vardakou, et al., 2010). Further complicating characterization of this class of abused
cannabinoids is the fact that their affinities for novel cannabinoid or noncannabinoid
NIH-PA Author Manuscript

receptors and the role that these receptors may play in mediating or modulating their
pharmacological effects are largely unknown. For example, the indole structure of many of
these compounds suggests that they may interact with one or more serotonin receptors.
Given that the focus of extant studies has been examination of the effects of these
compounds in assays specifically designed to detect cannabinoid activity, perhaps the most
pressing future research need is determination of the extent to which the pharmacology of
these compounds may differ from those of classical cannabinoids.

Acknowledgments
Preparation of this manuscript was supported by National Institute on Drug Abuse (NIDA) grants DA-031988 and
DA-03672; NIDA had no further role in the writing of the review or in the decision to submit the paper for
publication.

References
Aung MM, Griffin G, Huffman JW, Wu M, Keel C, Yang B, Showalter VM, Abood ME, Martin BR.
Influence of the N-1 alkyl chain length of cannabimimetic indoles upon CB(1) and CB(2) receptor
binding. Drug Alcohol Depend. 2000; 60:133–40. [PubMed: 10940540]
NIH-PA Author Manuscript

Breivogel CS, Childers SR. Cannabinoid agonist signal transduction in rat brain: comparison of
cannabinoid agonists in receptor binding, G-protein activation, and adenylyl cyclase inhibition. J
Pharmacol Exp Ther. 2000; 295:328–36. [PubMed: 10991998]
Brents LK, Gallus-Zawada A, Radominska-Pandya A, Vasiljevik T, Prisinzano TE, Fantegrossi WE,
Moran JH, Prather PL. Monohydroxylated metabolites of the K2 synthetic cannabinoid JWH-073
retain intermediate to high cannabinoid 1 receptor (CB1R) affinity and exhibit neutral antagonist to
partial agonist activity. Biochem Pharmacol. 2012; 83:952–61. [PubMed: 22266354]
Compton DR, Aceto MD, Lowe J, Martin BR. In vivo characterization of a specific cannabinoid
receptor antagonist (SR141716A): inhibition of delta 9-tetrahydrocannabinol-induced responses and
apparent agonist activity. J Pharmacol Exp Ther. 1996; 277:586–94. [PubMed: 8627535]
Compton DR, Gold LH, Ward SJ, Balster RL, Martin BR. Aminoalkylindole analogs: cannabimimetic
activity of a class of compounds structurally distinct from delta 9-tetrahydrocannabinol. J
Pharmacol Exp Ther. 1992; 263:1118–26. [PubMed: 1335057]
Compton DR, Rice KC, De Costa BR, Razdan RK, Melvin LS, Johnson MR, Martin BR. Cannabinoid
structure-activity relationships: Correlation of receptor binding and in vivo activities. J Pharmacol
Exp Ther. 1993; 265:218–26. [PubMed: 8474008]
Compton DR, Little PJ, Martin BR, Gilman JW, Saha JK, Jorapur VS, Sard HP, Razdan RK. Synthesis
and pharmacological evaluation of amino, azido, and nitrogen mustard analogues of 10-substituted
NIH-PA Author Manuscript

cannabidiol and 11- or 12-substituted delta 8-tetrahydrocannabinol. J Med Chem. 1990; 33:1437–
43. [PubMed: 2158563]
Cox AO, Daw RC, Mason MD, Grabenauer M, Pande PG, Davis KH, Wiley JL, Stout PR, Thomas
BF, Huffman JW. Use of SPME-HS-GC-MS for the analysis of herbal products containing
synthetic cannabinoids. J Anal Toxicol. 2012; 36:293–302. [PubMed: 22582264]
Denooz R, Vanheugen JC, Frederich M, de Tullio P, Charlier C. Identification and structural
elucidation of four cannabimimetic compounds (RCS-4, AM-2201, JWH-203 and JWH-210) in
seized products. J Anal Toxicol. 2013; 37:56–63. [PubMed: 23339188]
Devane WA, Dysarz FA, Johnson MR, Melvin LS, Howlett AC. Determination and characterization of
a cannabinoid receptor in rat brain. Mol Pharmacol. 1988; 34:605–13. [PubMed: 2848184]
Devane WA, Hanus L, Breuer A, Pertwee RG, Stevenson LA, Griffin G, Gibson D, Mandelbaum A,
Etinger A, Mechoulam R. Isolation and structure of a brain constituent that binds to the
cannabinoid receptor. Science. 1992; 258:1946–9. [PubMed: 1470919]

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 9

Dutta AK, Ryan W, Thomas BF, Singer M, Compton DR, Martin BR, Razdan RK. Synthesis,
pharmacology, and molecular modeling of novel 4-alkyloxy indole derivatives related to
cannabimimetic aminoalkyl indoles (AAIs). Bioorg Med Chem. 1997; 5:1591–600. [PubMed:
NIH-PA Author Manuscript

9313864]
Frost JM, Dart MJ, Tietje KR, Garrison TR, Grayson GK, Daza AV, El-Kouhen OF, Yao BB, Hsieh
GC, Pai M, Zhu CZ, Chandran P, Meyer MD. Indol-3-ylcycloalkyl ketones: effects of N1
substituted indole side chain variations on CB(2) cannabinoid receptor activity. J Med Chem.
2010; 53:295–315. [PubMed: 19921781]
Frost JM, Dart MJ, Tietje KR, Garrison TR, Grayson GK, Daza AV, El-Kouhen OF, Miller LN, Li L,
Yao BB, Hsieh GC, Pai M, Zhu CZ, Chandran P, Meyer MD. Indol-3-yl-tetramethylcyclopropyl
ketones: effects of indole ring substitution on CB2 cannabinoid receptor activity. J Med Chem.
2008; 51:1904–12. [PubMed: 18311894]
Gaoni Y, Mechoulam R. Isolation, structure, and partial synthesis of an active constituent of hashish. J
Amer Chem Soc. 1964; 86:1646–7.
Ginsburg BC, Schulze DR, Hruba L, McMahon LR. JWH-018 and JWH-073: Delta-
tetrahydrocannabinol-like discriminative stimulus effects in monkeys. J Pharmacol Exp Ther.
2012; 340:37–45. [PubMed: 21965552]
Grabenauer M, Krol WL, Wiley JL, Thomas BF. Analysis of synthetic cannabinoids using high-
resolution mass spectrometry and mass defect filtering: implications for nontargeted screening of
designer drugs. Anal Chem. 2012; 84:5574–81. [PubMed: 22724537]
Hanus L, Abu-Lafi S, Fride E, Breuer A, Vogel Z, Shalev DE, Kustanovich I, Mechoulam R. 2-
Arachidonyl glyceryl ether, an endogenous agonist of the cannabinoid CB1 receptor. Proc Natl
NIH-PA Author Manuscript

Acad Sci USA. 2001; 98:3662–5. [PubMed: 11259648]


Haubrich DR, Ward SJ, Baizman E, Bell MR, Bradford J, Ferrari R, Miller M, Perrone M, Pierson
AK, Saelens JK, et al. Pharmacology of pravadoline: a new analgesic agent. J Pharmacol Exp
Ther. 1990; 255:511–22. [PubMed: 2243340]
Huffman JW. Cannabimimetic indoles, pyrroles and indenes. Curr Med Chem. 1999; 6:705–20.
[PubMed: 10469887]
Huffman, JW. Cannabimimetic indoles, pyrroles, and indenes: Structure-activity relationships and
receptor interactions. In: Reggio, PH., editor. The cannabinoid receptors. New York: Humana
Press; 2009. p. 49-94.
Huffman JW, Padgett LW. Recent developments in the medicinal chemistry of cannabimimetic
indoles, pyrroles and indenes. Curr Med Chem. 2005; 12:1395–411. [PubMed: 15974991]
Huffman JW, Dai D, Martin BR, Compton DR. Design, synthesis and pharmacology of
cannabimimetic indoles. Bioorg Med Chem Lett. 1994; 4:563–6.
Huffman JW, Padgett LW, Isherwood ML, Wiley JL, Martin BR. 1-Alkyl-2-aryl-4-(1-
naphthoyl)pyrroles: New high affinity ligands for the cannabinoid CB(1) and CB(2) receptors.
Bioorg Med Chem Lett. 2006; 16:5432–5. [PubMed: 16889960]
Huffman JW, Mabon R, Wu MJ, Lu J, Hart R, Hurst DP, Reggio PH, Wiley JL, Martin BR. 3-
Indolyl-1-naphthylmethanes: new cannabimimetic indoles provide evidence for aromatic stacking
NIH-PA Author Manuscript

interactions with the CB(1) cannabinoid receptor. Bioorg Med Chem. 2003; 11:539–49. [PubMed:
12538019]
Huffman JW, Szklennik PV, Almond A, Bushell K, Selley DE, He H, Cassidy MP, Wiley JL, Martin
BR. 1-Pentyl-3-phenylacetylindoles, a new class of cannabimimetic indoles. Bioorg Med Chem
Lett. 2005a; 15:4110–3. [PubMed: 16005223]
Huffman JW, Zengin G, Wu MJ, Lu J, Hynd G, Bushell K, Thompson AL, Bushell S, Tartal C, Hurst
DP, Reggio PH, Selley DE, Cassidy MP, Wiley JL, Martin BR. Structure-activity relationships for
1-alkyl-3-(1-naphthoyl)indoles at the cannabinoid CB(1) and CB(2) receptors: steric and electronic
effects of naphthoyl substituents. New highly selective CB(2) receptor agonists. Bioorg Med
Chem. 2005b; 13:89–112. [PubMed: 15582455]
Järbe TU, Deng H, Vadivel SK, Makriyannis A. Cannabinergic aminoalkylindoles, including
AM678=JWH018 found in ‘Spice’, examined using drug (Delta9-tetrahydrocannabinol)
discrimination for rats. Behav Pharmacol. 2011; 22:498–507. [PubMed: 21836461]

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 10

Kavanagh P, Grigoryev A, Savchuk S, Mikhura I, Formanovsky A. UR-144 in products sold via the
Internet: Identification of related compounds and characterization of pyrolysis products. Drug Test
Anal. 2013 Epub ahead of print. 10.1002/dta.1456
NIH-PA Author Manuscript

Lainton JAH, Huffman JW, Martin BR, Compton DR. 1-Alkyl-3-(1-naphthoyl)pyrroles: A new
cannabinoid class. Tetrahedron Lett. 1995; 36:1401–4.
Lindigkeit R, Boehme A, Eiserloh I, Luebbecke M, Wiggermann M, Ernst L, Beuerle T. Spice: a never
ending story? Forensic Sci Int. 2009; 191:58–63. [PubMed: 19589652]
Little PJ, Compton DR, Johnson MR, Melvin LS, Martin BR. Pharmacology and stereoselectivity of
structurally novel cannabinoids in mice. J Pharmacol Exp Ther. 1988; 247:1046–51. [PubMed:
2849657]
Logan BK, Reinhold LE, Xu A, Diamond FX. Identification of synthetic cannabinoids in herbal
incense blends in the United States. J Forensic Sci. 2012; 57:1168–80. [PubMed: 22834927]
Manera C, Tuccinardi T, Martinelli A. Indoles and related compounds as cannabinoid ligands. Mini
Rev Med Chem. 2008; 8:370–87. [PubMed: 18473928]
Martin BR, Balster RL, Razdan RK, Harris LS, Dewey WL. Behavioral comparisons of the
stereoisomers of tetrahydrocannabinols. Life Sci. 1981; 29:565–74. [PubMed: 6268916]
Martin BR, Compton DR, Thomas BF, Prescott WR, Little PJ, Razdan RK, Johnson MR, Melvin LS,
Mechoulam R, Ward SJ. Behavioral, biochemical, and molecular modeling evaluations of
cannabinoid analogs. Pharmacol Biochem Behav. 1991; 40:471–8. [PubMed: 1666911]
Matsuda LA, Lolait SJ, Brownstein MJ, Young AC, Bonner TI. Structure of a cannabinoid receptor
and functional expression of the cloned cDNA. Nature. 1990; 346:561–4. [PubMed: 2165569]
NIH-PA Author Manuscript

Mechoulam R, Ben-Shabat S, Hanus L, Ligumsky M, Kaminski NE, Schatz AR, Gopher A, Almog S,
Martin BR, Compton DR, et al. Identification of an endogenous 2-monoglyceride, present in
canine gut, that binds to cannabinoid receptors. Biochem Pharmacol. 1995; 50:83–90. [PubMed:
7605349]
Munro S, Thomas KL, Abu-Shaar M. Molecular characterization of a peripheral receptor for
cannabinoids. Nature. 1993; 365:61–4. [PubMed: 7689702]
National Research Council. Guide for the Care and Use of Laboratory Animals. Washington, D.C:
National Academy Press; 1996.
Reggio PH, Basu-Dutt S, Barnett-Norris J, Castro MT, Hurst DP, Seltzman HH, Roche MJ, Gilliam
AF, Thomas BF, Stevenson LA, Pertwee RG, Abood ME. The bioactive conformation of
aminoalkylindoles at the cannabinoid CB1 and CB2 receptors: insights gained from (E)- and (Z)-
naphthylidene indenes. J Med Chem. 1998; 41:5177–87. [PubMed: 9857088]
Rinaldi-Carmona M, Barth F, Héaulme M, Shire D, Calandra B, Congy C, Martinez S, Maruani J,
Néliat G, Caput D, Ferrara P, Soubrié P, Brelière JC, Le Fur G. SR141716A, a potent and selective
antagonist of the brain cannabinoid receptor. FEBS Lett. 1994; 350:240–4. [PubMed: 8070571]
Rinaldi-Carmona M, Barth F, Millan J, Defrocq J, Casellas P, Congy C, Oustric D, Sarran M,
Bouaboula M, Calandra B, Portier M, Shire D, Breliere J, Le Fur G. SR 144528, the first potent
and selective antagonist of the CB2 cannabinoid receptor. J Pharmacol Exp Ther. 1998; 284:644–
50. [PubMed: 9454810]
NIH-PA Author Manuscript

Sañudo-Peña MC, Romero J, Seale GE, Fernandez-Ruiz JJ, Walker JM. Activational role of
cannabinoids on movement. Eur J Pharmacol. 2000; 391:269–74. [PubMed: 10729368]
Showalter VM, Compton DR, Martin BR, Abood ME. Evaluation of binding in a transfected cell line
expressing a peripheral cannabinoid receptor (CB2): identification of cannabinoid receptor subtype
selective ligands. J Pharmacol Exp Ther. 1996; 278:989–99. [PubMed: 8819477]
Thomas BF, Compton DR, Martin BR, Semus SF. Modeling the cannabinoid receptor: a three-
dimensional quantitative structure-activity analysis. Mol Pharmacol. 1991; 40:656–65. [PubMed:
1944237]
Uchiyama N, Kawamura M, Kikura-Hanajiri R, Goda Y. URB-754: A new class of designer drug and
12 synthetic cannabinoids detected in illegal products. Forensic Sci Int. 2013; 227:21–32.
[PubMed: 23063179]
Vardakou I, Pistos C, Spiliopoulou C. Spice drugs as a new trend: mode of action, identification and
legislation. Toxicol Lett. 2010; 197:157–62. [PubMed: 20566335]

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 11

Wiebelhaus JM, Poklis JL, Poklis A, Vann RE, Lichtman AH, Wise LE. Inhalation exposure to smoke
from synthetic “marijuana” produces potent cannabimimetic effects in mice. Drug Alcohol
Depend. 2012; 126:316–23. [PubMed: 22776442]
NIH-PA Author Manuscript

Wiley JL, Martin BR. Cannabinoid pharmacological properties common to other centrally acting
drugs. Eur J Pharmacol. 2003; 471:185–93. [PubMed: 12826237]
Wiley JL, Evans RL, Grainger DB, Nicholson KL. Age-dependent differences in sensitivity and
sensitization to cannabinoids and ‘club drugs’ in male adolescent and adult rats. Addict Biol. 2008;
13:277–86. [PubMed: 17850418]
Wiley JL, Marusich JA, Lefever TW, Grabenauer M, Moore KN, Thomas BF. Cannabinoids in
disguise: Δ9-Tetrahydrocannabinol-like effects of tetramethylcyclopropyl ketone indoles.
Neuropharmacology. 2013; 75:145–54. [PubMed: 23916483]
Wiley JL, Marusich JA, Martin BR, Huffman JW. 1-Pentyl-3-phenylacetylindoles and JWH-018 share
in vivo cannabinoid profiles in mice. Drug Alcohol Depend. 2012a; 123:148–53. [PubMed:
22127210]
Wiley JL, Smith VJ, Chen J, Martin BR, Huffman JW. Synthesis and pharmacology of 1-alkyl-3-(1-
naphthoyl)indoles: steric and electronic effects of 4- and 8-halogenated naphthoyl substituents.
Bioorg Med Chem. 2012b; 20:2067–81. [PubMed: 22341572]
Wiley JL, Compton DR, Dai D, Lainton JA, Phillips M, Huffman JW, Martin BR. Structure-activity
relationships of indole- and pyrrole-derived cannabinoids. J Pharmacol Exp Ther. 1998; 285:995–
1004. [PubMed: 9618400]
NIH-PA Author Manuscript
NIH-PA Author Manuscript

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 12
NIH-PA Author Manuscript

Figure 1.
Historical 3-point attachment model used to develop JWH-018. An overlay of
WIN55,212-2, Δ9-THC and JWH-018 was hypothesized, in which each cannabinoid would
NIH-PA Author Manuscript

attach to the CB1 receptor at the 3 locations specified in the figure.


NIH-PA Author Manuscript

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 13
NIH-PA Author Manuscript
NIH-PA Author Manuscript

Figure 2.
Effects of WIN55,212-2 on spontaneous activity (panel A), antinociception (panel B), rectal
temperature (panel C), and catalepsy (panel D) in adult male ICR mice. Spontaneous activity
was measured as total number of photocell beam interruptions during a 10-min session and
is shown as % inhibition of activity of the vehicle group. Antinociception in a tail flick assay
is expressed as the percent maximum possible effect (MPE) using a 10-s maximum test
latency as follows: [(test−control)/(10−control)]×100. Rectal temperature values are
expressed as the difference between control temperature (before injection) and temperature
following drug administration (Δ°C). During assessment for catalepsy, the total amount of
time (in s) that the mouse remained motionless on the ring apparatus (except for breathing
and whisker movement) was measured and was used as an indication of catalepsy-like
behavior. This value was divided by 300 s and multiplied by 100 to obtain percent
immobility, as shown in the figure. Values represent the mean (± SEM) of 5–6 mice per
group. The typical maximal effect observed for each measure is shown in the box at the
upper left of each panel. ED50s for data presented in these panels are provided in Table 1.
NIH-PA Author Manuscript

Life Sci. Author manuscript; available in PMC 2015 February 27.


Wiley et al. Page 14
NIH-PA Author Manuscript

Figure 3.
Templates for major structural classes of synthetic cannabinoids discussed in this review:
naphthoylindoles, naphthoylpyrroles, 1-pentyl-3-phenylacetylindoles and
NIH-PA Author Manuscript

tetramethylcyclopropyl ketone indoles. Additional structural templates are shown in the


tables.
NIH-PA Author Manuscript

Life Sci. Author manuscript; available in PMC 2015 February 27.


NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Table 1
Cannabinoid receptor binding and in vivo effects of indoles with structural variations in the side chain and indole substituent
Wiley et al.

Affinities (nM) In Vivo Tests*


Compound R R′
CB1 CB2 SA %MPE RT RI

87% (10) 100% (10) −5.0 (10) 87% (10)


WIN55,212-2** morpholino H 1.9 ± 1a 0.28 ± 0.16a
1.3 0.3 0.6 2.2

JWH-446 propyl Br 3322 ± 430 2011 ± 239 83% (30) 76% (30) −3.6 (30) 64% (30)
JWH-443 butyl Br 364 ± 14 332 ± 84 ND ND ND ND
JWH-444 pentyl Br 145 ± 8 207 ± 5 79% (30) 100% (30) −5.0 (30) 76% (30)
JWH-445 hexyl Br 197 ± 14 499 ± 117 ND ND ND ND

JWH-453 propyl I 1967 ± 478 778 ± 50 inactive (30) inactive (30) inactive (30) inactive (30)

Life Sci. Author manuscript; available in PMC 2015 February 27.


JWH-455 butyl I 686 ± 48 722 ± 157 79% (30) 68% (30) −3.3 (30) inactive (30)
JWH-452 pentyl I 178 ± 6 193 ± 19 74% (30) 62% (30) −5.4 (30) 45% (30)
JWH-454 hexyl I 152 ± 10 364 ± 15 66% (30) inactive (30) −2.5 (30) ND

*
SA = spontaneous activity, %MPE = % maximum possible antinociceptive effect, RT = rectal temperature, RI = ring immobility. ND = not determined. Single dose tests are indicated by magnitude of
effect and dose tested (mg/kg, in parentheses). “Inactive” refers to doses that produced less than 30% of the maximal cannabimimetic effect for the measure (see Figure 2).
**
Values for in vivo tests for WIN55,212-2 were calculated based on data shown in Figure 2 and are presented for comparison and for illustration of the system used for data presented in all tables.
Numbers in top row exemplify testing of a single 10 mg/kg dose; values on the bottom row show ED50s for dose-effect data presented in Figure 2.

a
Showalter, et al., 1996
Page 15
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Table 2
Cannabinoid receptor binding and in vivo effects of 2-aryl-4-(1-naphthoyl)-N-pentyl-pyrroles
Wiley et al.

Affinities (nM) a In VivoTests*


Compound R
CB1 CB2 SA %MPE RT RI

JWH-309 1-naphthyl 41 ± 3 49 ± 7 inactive (3) inactive (3) inactive (3) ND


JWH-366 3-pyridyl 191 ± 12 24 ± 1 91% (30) 86% (30) −4.25 (30) 44% (30)

Life Sci. Author manuscript; available in PMC 2015 February 27.


JWH-370 2-methylphenyl 5.6 ± 0.4 4.0 ± 0.5 93% (10) 100% (10) −6.2 (10) 94% (10)
JWH-346 3-methylphenyl 67 ± 6 39 ± 2 1.5 1.3 2.6 ND

JWH-365 2-ethylphenyl 17 ± 1 3.4 ± 0.2 86% (10) 95% (10) −5.1 (10) 48% (10)
JWH-364 4-ethylphenyl 34 ± 3 29 ± 1 inactive (30) 76% (30) −5.1 (10) 48% (10)

JWH-371 4-butylphenyl 42 ± 1 64 ± 2 76% (30) 74% (30) −2.7 (30) inactive (30)
JWH-367 3-methoxyphenyl 53 ± 2 23 ± 1 83% (30) 100% (30) −5.0 (30) 58% (30)

JWH-307 2-fluorophenyl 7.7 ± 1.8 3.3 ± 0.2 0.42 0.53 0.96 ND


Page 16
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Wiley et al.

Affinities (nM) a In VivoTests*


Compound R
CB1 CB2 SA %MPE RT RI

JWH-308 4-fluorophenyl 41± 1 33 ± 2 2.4 60% (3) −2.6 (3) ND

JWH-372 2-trifluorophenyl 77 ± 2 8.2 ± 0.2 63% (30) 100% (30) −4.6 (30) 27% (30)
JWH-363 3-trifluorophenyl 245 ± 5 71 ± 1 48% (30) inactive (30) −2.0 (30) inactive (30)

Life Sci. Author manuscript; available in PMC 2015 February 27.


SA = spontaneous activity, %MPE = % maximum possible antinociceptive effect, RT = rectal temperature, RI = ring immobility. ND = not determined. Single dose tests are indicated by magnitude of
effect and dose tested (mg/kg, in parentheses). “Inactive” refers to doses that produced less than 30% of the maximal cannabimimetic effect for the measure (see Figure 2). Whenever supplies allowed
determination of a dose-effect curve, ED50s (mg/kg) are provided for active compounds.

a
Huffman, et al., 2006
Page 17
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Table 3
Cannabinoid receptor binding and in vivo effects of methoxy naphthoyl indoles
Wiley et al.

Substituents Affinities (nM) In Vivo Tests*


Compound
R R′ X** CB1 CB2 SA %MPE RT RI

JWH-391 butyl H 2-OCH3 > 10,000 1236 ± 103 inactive (30) 67% (30) inactive (30) inactive (30)

JWH-390 butyl CH3 2-OCH3 > 10,000 > 10,000 stim (30) inactive (30) inactive (30) inactive (30)

JWH-388 pentyl H 2-OCH3 2398 ± 228 147 ± 8 inactive (30) inactive (30) inactive (30) inactive (30)

JWH-389 pentyl CH3 2-OCH3 > 10,000 > 10,000 inactive (30) inactive (30) −2.1 (30) inactive (30)

JWH-080 butyl H 4-OCH3 5.6 ± 1a 2.2 ± 1.3a 2.7 1.5 0.88 ND

JWH-081 pentyl H 4-OCH3 1.2 ± 0.03a 12.5 ± 2.23 a 0.06 0.084 0.06 ND

JWH-082 hexyl H 4-OCH3 5.3 ± 0.80a 6.4 ± 0.94a 1.7 0.62 1.1 ND

JWH-083 heptyl H 4-OCH3 106 ± 12a 102 ± 50a stim 24 28 ND

Life Sci. Author manuscript; available in PMC 2015 February 27.


JWH-411 butyl H 6-OCH3 2036 ± 203 561 ± 54 inactive (30) inactive (30) inactive (30) inactive (30)

JWH-410 butyl CH3 6-OCH3 2682 ± 397 322 ± 46 64% (30) 75% (30) −2.7 (30) 33% (30)

JWH-408 pentyl H 6-OCH3 190 ± 12 91 ± 3 80% (30) 96% (30) −4.3 (30) 62% (30)

JWH-409 pentyl CH3 6-OCH3 746 ± 77 232 ± 20 83% (30) 42% (30) −3.9 (30) 38% (30)

*
SA = spontaneous activity, %MPE = % maximum possible antinociceptive effect, RT = rectal temperature, RI = ring immobility. ND = not determined. Stim = stimulation of locomotor activity of at least
10% above vehicle levels (i.e., −10% inhibition of activity). Single dose tests are indicated by magnitude of effect and dose tested (mg/kg, in parentheses). “Inactive” refers to doses that produced less than
30% of the maximal cannabimimetic effect for the measure (see Figure 2). Whenever supplies allowed determination of a dose-effect curve, ED50s (mg/kg) are provided for active compounds.

**
X indicates substituent and its position on the naphthoyl group.
a
Aung, et al., 2000
Page 18
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript

Table 4
Cannabinoid receptor binding and in vivo effects of miscellaneous indoles

Affinities (nM) In Vivo Tests*


Compound Structure
Wiley et al.

CB1 CB2 SA %MPE RT RI

JWH-001 > 10,000 ND inactive (100) inactive (100) inactive (100) inactive (100)

JWH-002 > 10,000 ND inactive (100) inactive (100) inactive (100) ND

JWH-317 101 ± 1 257 ± 3 89% (30) 100% (30) −5.5 (30) ND

Life Sci. Author manuscript; available in PMC 2015 February 27.


*
SA = spontaneous activity, %MPE = % maximum possible antinociceptive effect, RT = rectal temperature, RI = ring immobility. ND = not determined. Single dose tests are indicated by magnitude of
effect and dose tested (mg/kg, in parentheses). “Inactive” refers to doses that produced less than 30% of the maximal cannabimimetic effect for the measure (see Figure 2).
Page 19

You might also like